Vous êtes sur la page 1sur 32

Measuring Performance

of Exchange Traded Funds∗


Marlène Hassine Thierry Roncalli
ETF Strategy Research & Development
Lyxor Asset Management, Paris Lyxor Asset Management, Paris
marlene.hassine@lyxor.com thierry.roncalli@lyxor.com
February 2013

Abstract
Fund selection is an important issue for investors. This topic has spawned abundant
academic literature. Nonetheless, most of the time, these works concern only active
management, whereas many investors, such as institutional investors, prefer to invest in
index funds. The tools developed in the case of active management are also not suitable
for evaluating the performance of these index funds. This explains why information
ratios are usually used to compare the performance of passive funds. However, we
show that this measure is not pertinent, especially when the tracking error volatility of
the index fund is small. The objective of an exchange traded fund (ETF) is precisely
to offer an investment vehicle that presents a very low tracking error compared to its
benchmark. In this paper, we propose a performance measure based on the value-at-risk
framework, which is perfectly adapted to passive management and ETFs. Depending
on three parameters (performance difference, tracking error volatility and liquidity
spread), this efficiency measure is easy to compute and may help investors in their
fund selection process. We provide some examples, and show how liquidity is more of
an issue for institutional investors than retail investors.
Keywords: Passive management, index fund, ETF, information ratio, tracking error, liq-
uidity, spread, value-at-risk.
JEL classification: G11.

1 Introduction
The market portfolio concept has a long history and dates back to the seminal work of
Markowitz (1952). In that paper, Markowitz defines precisely what portfolio selection means:
“the investor does (or should) consider expected return a desirable thing and variance of re-
turn an undesirable thing”. Indeed, Markowitz shows that an efficient portfolio is a portfolio
that maximizes the expected return for a given level of risk (corresponding to the variance
of return). Markowitz concludes that there is not only one optimal portfolio, but a set of
optimal portfolios called the efficient frontier (represented by the solid blue curve in Figure
1). By studying liquidity preference, Tobin (1958) shows that the efficient frontier becomes a
∗ We are profoundly grateful to Bou Ly Wu for his support with data management and the computation

of liquidity spreads on limit order books. We would also like to thank Arnaud Llinas, Raphaël Dieterlen,
Valérie Lalonde, François Millet and Matthieu Mouly for their helpful comments.

1
Measuring Performance of Exchange Traded Funds

straight line in the presence of a risk-free asset. If we consider a combination of an optimized


portfolio and the risk-free asset, we obtain a straight line (represented by the dashed black
line in Figure 1). But one straight line dominates all the other straight line and the efficient
frontier. It is called the Capital Market Line (CML), which corresponds to the green dashed
line in Figure 1. In this case, optimal portfolios correspond to a combination of the risk-
free asset and one particular efficient portfolio named the tangency portfolio. Sharpe (1964)
summarizes the results of Markowitz and Tobin as follows: “the process of investment choice
can be broken down into two phases: first, the choice of a unique optimum combination of
risky assets; and second, a separate choice concerning the allocation of funds between such
a combination and a single riskless asset”. This two-step procedure is today known as the
Separation Theorem.

One of the difficulties faced when computing the tangency portfolio is that of precisely
defining the vector of expected returns of the risky assets and the corresponding covari-
ance matrix of returns. In 1964, Sharpe developed the CAPM theory and highlighted the
relationship between the risk premium of the asset (the difference between the expected
return and the risk-free rate) and its beta (the systematic risk with respect to the tangency
portfolio). Assuming that the market is at equilibrium, he showed that the prices of assets
are such that the tangency portfolio is the market portfolio, which is composed of all risky
assets in proportion to their market capitalization. That is why we use the terms, tangency
portfolio and market portfolio indiscriminately nowadays.

Figure 1: Efficient frontier and the tangency portfolio

Another step forward was made by Jensen (1968), who measured the performance of 115
(equity) mutual funds using the alpha measure and concluded that:
“The evidence on mutual fund performance indicates not only that these 115
mutual funds were on average not able to predict security prices well enough to

2
Measuring Performance of Exchange Traded Funds

outperform a buy-the-market-and-hold policy, but also that there is very little


evidence that any individual fund was able to do significantly better than that
which we expected from mere random chance”.

The seminal work of Michael Jensen is the starting point of the development of passive
management. Indeed, the first index fund was launched in 1971 by John McQuown at Wells
Fargo Bank for the Samsonite luggage company (Bernstein, 1992). It took a long time for
investors to accept the ideas of Markowitz, Sharpe and Jensen, but passive management
represents now a large part of the asset management industry as shown in Figure 2.

Figure 2: Index funds’ share in the equity mutual fund market


18%

16%

14%

12%

10%

8%

6%

4%

2%

0%
1997 1998 1999 2000 2001 2002 2003 2004 2005 2006 2007 2008 2009 2010 2011

There is an abundance of literature that compares the performance of active managers.


Indeed, academics have developed many econometrics tools in this domain to measure their
alpha and their skill (Grinblatt and Titman, 1989; Blake et al., 1993; Carhart, 1997; Barras
et al., 2010). Professionals have also created databases and ranking systems to evaluate
mutual funds. For instance, Morningstar is certainly one of the best known scoring platforms.
The literature is scarcer when it comes to measuring the performance of passive management.
Generally, one uses the same quantitative tools as those developed for passive management
to investigate the efficiency of an index fund. The aim of this paper is to provide more
appropriate tools for investor looking to invest in index funds, and in exchange traded funds
in particular.

The paper is organized as follows. Section 2 presents the efficiency measure, which is in
fact a risk measure. In particular, we describe the reasoning behind this indicator and its
rationale. Empirical results are reported in Section 3. In section 4, we consider different
variations on the tracker efficiency measure. For example, we consider other definitions of
the risk measure and the liquidity spread. Finally, in Section 5 we draw our conclusions

3
Measuring Performance of Exchange Traded Funds

2 Measuring the efficiency of exchange traded funds


2.1 Performance or efficiency measurement?
2.1.1 Understanding fund picking in active management
Fund picking could be summarized as a two-step procedures:

• Defining a universe of funds, that are sensitive to certain risk factors;

• Picking one or more elements of this universe by combining quantitative and qualitative
criteria.

The first step is crucial in order to define a universe of mutual funds, that are sufficiently
homogeneous in terms of risk analysis. For instance, if we want to invest in equities, it
could be useful to describe the universe more accurately. Does the investment concern
global equities, regional equities or country equities? Does the investment definition relate to
specific sectors, by focusing on or excluding some of them? What is the bias of the investment
universe in terms of style analysis? The capacity of the investor to clearly define the borders
of the universe is the most important part of the fund picking process, because the second
step is relatively straightforward once the first step has been accurately completed. For
example, investing in large capitalization equities located in the Eurozone that do not belong
to the banking system, and are recognized as socially responsible, effectively reduces the
universe to a small number of mutual funds.

Nevertheless, defining the investment universe is not always straightforward, for many
reasons. First, the category may be very large with numerous mutual funds. Suppose we
wanted to invest in sovereign bonds in the Eurozone. According to the Morningstar database
there are more than 200 such funds available. Second, a category may not necessarily exist
if the criteria are too specific1 . Third, investment styles are heterogeneous, because they
cannot be defined in a precise way. Take value and growth styles, for example. Everybody
knows what they mean, but if we ask two people to classify the components of the S&P 500
index according to growth/value risk factors, we will get two different answers. This shows
that investing in active management may have a significant cost, because thorough research
is necessary to find accurate information. That is why, in the end, many investments are
based on past performance.

2.1.2 Why is fund picking different with passive management?


With passive management, fund picking is more straightforward. Indeed, the fund universe
is clearly defined once the benchmark has been chosen and the problem of performance
attribution between alpha and beta components is not relevant. The investor’s goal is also
to identify the investment vehicle that tracks the chosen index most accurately. Rating
systems such as Morningstar and Europerformance (see Box 1) are thus not suited to this
selection challenge, because they are more concerned by absolute performance. For passive
managers, absolute performance is meaningless. In an ideal world, they would like to buy
and sell the investment at any moment and have exactly the same return as the index.
Passive managers thus focus on factors other than absolute return:
1 For example, an investor seeking exposure to investment grade sovereign bonds in the Eurozone, exclud-

ing German bonds (because the yield is too low) and Spanish bonds (because they are too risky), would
struggle to find mutual funds that satisfy these criteria.

4
Measuring Performance of Exchange Traded Funds
' $

Box 1: The Europerformance/Edhec rating methodology

The ratings are constructed by combining three criteria. First, one measures extreme
risks using a Cornish-Fisher value-at-risk at a 99% confidence level. If the VaR of the
fund is too high, the fund is not rated. Second, the alpha of the fund is estimated. It is
done in two steps. In a first step, the style analysis of Sharpe is considered to build the
composite benchmark of the fund. When this benchmark is selected, one deduces alpha
by a regression analysis:

n ( )
(j)
Rt − r = α + βj Rt − r + εt
j=1

(j)
where Rt is the return of the fund, Rt is the return of the ith selected benchmark and
r is the risk-free rate. The alpha criterion allows us to distinguish funds rated below and
above 3 stars. For example, the 4 and 5 stars correspond to funds which have a strictly
positive alpha. The distinction between 4 and 5 stars is done using a gain frequency
measure. It corresponds to the number of times in percent that the fund has delivered
performance that was better than that of its benchmark. If the gain frequency measure
is less than 50%, the fund is rated 4 stars, otherwise it is rated 5 stars. Finally, a super
rating is attributed to funds rated 5 stars and which have a Hurst exponent H larger
than 1/2.
& %

1. Management fees, other costs and additional revenues;


2. Tracking error volatility;
3. Liquidity;
4. Structuration risks.

The first factor is important because it impacts on the net performance of the invest-
ment vehicle. For instance, high management fees generally have a negative impact on the
performance of mutual funds. Indeed, Elton et al. (1993) investigated the informational
efficiency of mutual fund performance for the period 1964-1984 and concluded:
“There is statistically significant relationship between alpha and expenses and
a relationship that can clearly be seen by examining the quintiles. Higher ex-
penses are associated with poorer performance. Management does not increase
performance by an amount sufficient that justify higher fees”.
The results are the same for index funds (Elton et al., 2004). Nevertheless, management
fees are not the only factor that impact directly on the return of an index fund. We can cite
for example brokerage costs, dividend taxes, etc. A fund may also benefit from revenues
derived from securities lending, fiscal optimization or index arbitrage2 .

The second factor measures the relative risk of the fund with respect to the index. It
corresponds in general to the volatility of the tracking error, which is defined as the difference
2 Index arbitrage concerns addition/deletion mechanisms of the index and corporate actions as subscrip-

tion rights or optional dividend management (Dieterlen and Hereil, 2012).

5
Measuring Performance of Exchange Traded Funds

between fund and index returns. It indicates how the tracker performance may deviate from
the index performance. In the words of Harry Markowitz, it is ‘an undesirable thing’ for
the investor. This performance difference comes from the fact that replication strategies
cannot be perfect, because the invested portfolio can not track the index weights exactly
all the time. We generally distinguish between three replication methods: full replication,
optimized sampling and synthetic replication. In full replication, the fund manager invests in
all index securities in the same proportion as the underlying index. But the portfolio weights
can never be identical to the theoretical index weights, unless the fund manager buys the
whole market3 . Moreover, the portfolio weights are not static, as a result of subscriptions
and redemptions, dividends, cash drag, etc.

The third factor is a key element for passive funds. The liquidity of the fund is inherently
connected to the liquidity of the index. A S&P 500 index fund is certainly more liquid than
an MSCI emerging markets index fund. But the investment universe is not the only factor
in fund liquidity. These funds are mainly bought by institutional investors, especially in
Europe. Liquidity is then crucial when they carry out subscriptions or redemptions. These
decisions may have a big impact on the market. Moreover, the fund manager has little
margin for discretionary decisions, because he has to follow the systematic index strategy.
The index ignores more and less liquidity, but the index fund does not. Some recent events,
like the March 2011 Tohoku earthquake in Japan or the Greek debt crisis, have shown
how difficult it can be to manage the liquidity of an index fund in the face of such stress
scenarios. Redemptions from Japanese equity index funds were difficult to execute in the
days after the catastrophe in northern Japan on Friday March 11, 2011. In the same way,
the decision of EuroMTS to keep Greek bonds in the EuroMTS Eurozone Govies index (ex
CNO) after Greece was downgraded from investment grade in April 2010 caused problems in
managing index funds that were benchmarked to this index. The problem is more complex
with trackers, because they offer intra-day liquidity. And this liquidity has a cost, which
can be measured by bid-ask spreads.

Another element that is important for investors is the operational structure of an ETF.
We generally distinguish two main structures (physical and synthetic) used in the European
markets. Even if the operational structure may be sometimes an important criteria for
investors, it is more a go/no go check test than a statistic that can be taken into account in
the design of an efficiency measure.

2.1.3 Portfolio optimization when there is a benchmark


We may think that the Markowitz approach is less relevant when there is a benchmark.
However, it could easily be adapted by replacing the volatility of the portfolio by the volatility
of the tracking error. Let b (or x) be the vector of weights of the benchmark (or the active
portfolio). We note n the number of assets in the investment universe and Ri the return of
3 Let us assume that the index is composed of two stocks. Let N
i,t and Pi,t be the number of outstanding
shares and the price of the ith stock at time t. The index weights are given by:
Ni,t Pi,t
wi,t = ∑2
i=1 Ni,t Pi,t

For example, if N1,t = 136017, N2,t = 87123, P1,t = 110.54, N2,t = 125.23, we have w1,t = 57.95% and
w2,t = 42.05%. If the value of the fund is $1000, full replication involves buying respectively 5 and 3 shares
of the first and second stocks, whereas the cash represents $71.61. The portfolio weights of the two stocks
are then w1,t = 55.27% and w2,t = 37.57%. At time t + 1, we assume that there is a subscription of $60. If
the prices remain the same, we could buy another share of either the first stock or the second stock. In the
two cases, we obtain portfolio weights that are different from the index weights.

6
Measuring Performance of Exchange Traded Funds

the asset i. The tracking error of the active portfolio x with respect to the benchmark is the
difference between the return of the portfolio R (x) and the return of the benchmark R (b):
e = R (x) − R (b)
∑n ∑
n
= xi Ri − bi Ri
i=1 i=1
⊤ ⊤
= x R−b R

= (x − b) R
The expected tracking error is defined as follows:
µ (x | b) = E [e]

= (x − b) µ
whereas tracking error volatility is equal to4 :
σ 2 (x | b) = σ 2 (e)

= (x − b) Σ (x − b)
The objective of the investor is to maximize the expected tracking error with a constraint
on the tracking error volatility:

x⋆ = arg max (x − b) µ
u.c. 1⊤ x = 1 and σ (e) ≤ σ ⋆
We gave an example of the tracking-error efficient frontier in Figure 3. It is a straight line
when there is no restriction (Roll, 1992). If we impose some constraints, the efficient frontier
is moved to the left and is no longer a straight line.

To compare the performance of different portfolios, a better measure than the Sharpe
ratio is the information ratio, which is defined as follows (Grinold and Kahn, 2000):
µ (x | b)
IR (x | b) =
σ (x | b)

(x − b) µ
= √

(x − b) Σ (x − b)
If we consider a combination of the benchmark b and the active portfolio x, the composition
of the portfolio is:
y = (1 − α) b + αx
with α ≥ 0 the proportion of wealth invested in the portfolio x. We get:

µ (y | b) = (y − b) µ = αµ (x | b)
and:

σ 2 (y | b) = (y − b) Σ (y − b) = α2 σ 2 (x | b)
We deduce that:
µ (y | b) = IR (x | b) · σ (y | b)
It is the equation of a linear function between the expected tracking error and the tracking
error volatility of the portfolio y. It implies that the efficient frontier is a straight line:
4 In IOSCO terminology, µ (x | b) corresponds to Tracking Difference (TD) whereas σ (x | b) is called

Tracking Error (TE).

7
Measuring Performance of Exchange Traded Funds

Figure 3: Tracking-error efficient frontier

“If the manager is measured solely in terms of excess-return performance, he or


she should pick a point on the upper part of this efficient frontier. For instance,
the manager may have a utility function that balances expected value added
against tracking-error volatility. Note that because the efficient set consists of
a straight line, the maximal Sharpe ratio is not a usable criterion for portfolio
allocation” (Jorion, 2003, page 172).

If we add some other constraints to the portfolio optimization problem, the efficient frontier
is no more a straight line. In this case, one optimized portfolio dominates all the other
portfolios. It is the portfolio that belongs to the efficient frontier and the straight line that
is tangent to the efficient frontier. It is also the portfolio that maximizes the information
ratio. An illustration is provided in Figure 4. In this case, we check that the tangency
portfolio is the one that maximizes the tangent θ, which is the information ratio:

BC
tan θ =
AB
µ (x | b)
=
σ (x | b)
= IR (x | b)

2.1.4 The information ratio as a selection criteria for benchmarked funds


To understand why the information ratio is suitable for comparing benchmarked funds, we
consider the example given in Figure 5. It is obvious that portfolio x2 is preferable to
portfolio x1 because it has a better excess-return performance for the same tracking-error
volatility. If we compare portfolios x3 and x2 , they don’t have the same tracking-error

8
Measuring Performance of Exchange Traded Funds

Figure 4: The geometry of the information ratio

volatility. Nevertheless, we could build a portfolio x4 by mixing benchmark b and portfolio


x2 such that x4 has the same tracking-error volatility as x3 . According to the previous
analysis, we get: {
x4 = (1 − α) b + αx2
α = σ (x3 | b) /σ (x2 | b)
In this case, portfolio x4 is preferable to portfolio x3 because it performs better. We can
conclude that x2 is preferable to x3 . The information ratio of x4 is by its very nature equal
to the information ratio of x2 . So, to select a portfolio, we simply compare their information
ratios. We thus get the following result5 .

Proposition 1 Let x and y be two portfolios benchmarked on the same index b. x is prefer-
able to y if and only if the information ratio of x is greater than the information ratio of
y:
x ≻ y ⇔ IR (x | b) ≥ IR (y | b)

2.1.5 Pitfalls of the information ratio


This proposition is the core of fund selection processes for benchmarked funds. However,
it is only valid for funds with significant tracking-error volatility. Indeed, to establish this
proposition, we have made the assumption that we can replicate the benchmark exactly. In
real life, we need to use a tracker to proxy the benchmark. Let y be the combination of
tracker x0 and portfolio x. We get:

y = (1 − α) x0 + αx
5 We assume that the information ratio is positive.

9
Measuring Performance of Exchange Traded Funds

Figure 5: Comparing benchmarked portfolios

In Appendix A.1, we show that:


(1 − α) µ (x0 | b) + αµ (x | b)
IR (y | b) = √ (1)
(1 − α) σ 2 (x0 | b) + ασ 2 (x | b) + (α2 − α) σ 2 (x | x0 )
If µ (x0 | b) = 0 and σ (x0 | b) = 0, we get the previous result:
IR (y | b) = IR (x | b)
In general, the tracker is such that6 µ (x0 | b) < 0 and σ (x0 | b) > 0. In this case, we obtain:
IR (y | b) < IR (x | b)
and these two information ratios are closed when the tracking-error volatility of the portfolio
x is larger than that of the tracker x0 – σ (x0 | b) ≪ σ (x | b). In other cases the use of the
information ratios may be not appropriate.

Let us illustrate this drawback with an example. We assume that µ (x0 | b) = −5 bps,
µ (x1 | b) = 45 bps, σ (x1 | b) = 45 bps, σ (x1 | x0 ) = 25 bps and σ (x0 | b) = 5 bps. Portfolio
x1 and tracker x0 are represented in Figure 6. Next we consider a second portfolio x2 with the
following characteristics µ (x2 | b) = 15 bps and σ (x2 | b) = 15 bps. We get IR (x1 | b) = 1.29
and IR (x2 | b) = 1.0. Using Proposition 1, we can deduce that x1 is better than x2 because
we could build the portfolio x3 that dominates x2 :
x3 ≻ x2 ⇒ x1 ≻ x2
But the problem is that the information ratio of x3 could not reached because the benchmark
could only be replicated by tracker x0 . In this case, the combination of x1 and x0 gives x4
6 Due to the management fees.

10
Measuring Performance of Exchange Traded Funds

and we find that x2 ≻ x4 . In this example, if we target a tracking-error volatility of 15 bps,


we conclude that x2 dominates x1 even if the information ratio of x1 is larger. We conclude
that:
IR (x | b) > IR (y | b) ; x ≻ y

This problem concerns benchmarked funds with low tracking-error volatility in particular.
In this case, the information ratio cannot be used to select funds. This is especially true
for trackers that aim to have the lowest tracking-error volatility. Moreover, the information
ratio is not appropriate for measuring the relative performance of the tracker that presents
the lowest tracking error. Indeed, it is generally a fund that has a small but negative excess-
return performance. Because its information ratio is negative, this tracker is not efficient
from the information ratio perspective. It means that no investors are interested in using
it. This theoretical point of view is in contradiction with practice, because passive investors
like to consider funds with the smallest tracking error volatility.

Figure 6: Information ratio with the tracker

Remark 1 Another problem arises from the fact that the information ratio ignores the
magnitude of tracking-error volatility. For instance, if we consider two trackers with the
following characteristics:

Tracker µ (x0 | b) σ (x0 | b) IR (x0 | b)


#1 0.02% 0.03% 0.66
#2 0.40% 0.50% 0.80

the second tracker is preferred to the first tracker if we use the information ratio as the
selection criteria. However, the first tracker does a better replication job than the second
one.

11
Measuring Performance of Exchange Traded Funds

2.2 A comprehensive efficiency indicator for trackers


Let us consider a simple model with two periods. The investor buy the tracker x at time
t = 0 and sells it at time t = 1. Note the corresponding tracking error e. The relative PnL
of the investor with respect to the benchmark b is:
Π (x | b) = e − s (x | b)
where s (x | b) is the bid-ask spread of the tracker. We define the loss L (x | b) of the investor
as follows:
L (x | b) = −Π (x | b)
The tracker efficiency measure is a risk measure applied to the loss function L (x | b) of the
investor. We propose to use value-at-risk, which is today commonly accepted as a standard
risk measure. In this case, the efficiency measure ζα (x | b) is defined as follows7 :
ζα (x | b) = − {inf ζ : Pr {L (x | b) ≤ ζ} ≥ α}
This means that the investor has a probability of 1 − α of losing an amount greater than
−ζα (x | b). Let F be the probability distribution function of L (x | b). We get:
ζα (x | b) = −F−1 (α)
If we consider the tracking error model of the previous section and if we assume that asset
returns are Gaussian, we obtain the closed-form formula shown in Appendix A.2. We can
then derive the following definition:
Definition 1 The efficiency measure ζα (x | b) of the tracker x with respect to the benchmark
b corresponds to:
ζα (x | b) = µ (x | b) − s (x | b) − Φ−1 (α) σ (x | b) (2)
where µ (x | b) is the expected value of the tracking error, s (x | b) is the bid-ask spread and
σ (x | b) is the volatility of the tracking error.
We illustrate the computation of the efficiency measure in Figure 7 when the confidence
level α is set to 95%. We assume that µ (x | b) = 50 bps, σ (x | b) = 40 bps and s (x | b) = 20
bps. The curve corresponds to the density function of the PnL of the investor. It is not
centered on µ (x | b) because of the cost of selling the tracker. It explains that the mean of
the P&L is 30 bps. According to the Gaussian hypothesis, there is a 50% probability that
the investor’s gain is greater than 30 bps, but there is also a 50% probability that investor’s
gain will be less than 30 bps. The probability of the investor making a loss is therefore equal
to 22.66%. To compute the efficiency measure of the tracker, we take the quantile at 5%.
We finally obtain ζα (x | b) = −35.79 bps.

In Figure 8, we show the impact of the different factors on the efficiency measure. The
solid line corresponds to the previous parameter values whereby we modify one parameter
value for each dashed line. For instance, in the first panel, we consider that µ (x | b) is
equal to 70 bps. This tracker is preferred as the original tracker because it has a better
excess-return performance, the same tracking error volatility and the same bid-ask spread.
In the second panel, the second tracker has a wider spread8 . In this case, we prefer the
first tracker. For the third panel, we obtain the same conclusion if the second tracker has a
greater tracking error volatility. We can then deduce the following result:
7 We consider the opposite of the loss in order to obtain an ascending order: the bigger the efficiency

measure, the better the tracker.


8 It is equal to 40 bps.

12
Measuring Performance of Exchange Traded Funds

Figure 7: Computing the efficiency measure

Figure 8: The effets of the different factors on the efficiency measure

13
Measuring Performance of Exchange Traded Funds

Proposition 2 Let x and y be two trackers benchmarked to the same index b. x is preferable
to y if and only if the efficiency measure of x is larger than the efficiency measure of y:

x ≻ y ⇔ ζα (x | b) ≥ ζα (y | b)

Example 1 We consider two trackers x and y with the following characteristics: µ (x | b) =


40 bps, s (x | b) = 20 bps, σ (x | b) = 30 bps, µ (x | b) = 30 bps, s (x | b) = 15 bps, σ (x | b) =
20 bps. We have Φ−1 (α) = 1.65 at the confidence level 95%. It follows that:

ζα (x | b) = 40 − 20 − 1.65 × 30
= −29.5

and:

ζα (y | b) = 30 − 15 − 1.65 × 20
= −18.0

Because ζα (y | b) > ζα (x | b), we can deduce that tracker y is more efficient than tracker x.
A loss of more than 29.5 bps has a 5% probability of occurring for tracker x. In the case of
tracker y, this loss is only 18 bps for the same level of probability.

3 Empirical results
In this paragraph, we apply the efficiency measure ζα (x | b) to the European ETF market.
The confidence level is set at 95%. Let [t0 , T ] be the study period with n trading dates. We
compute ζα (x | b) as follows:

ζα (x | b) = µ̂ (x | b) − ŝ (x | b) − 1.65 · σ̂ (x | b)

where µ̂ (x | b), ŝ (x | b) and σ̂ (x | b) are the estimated statistics for the given study period.
We have:
( )1/(T −t0 ) ( )1/(T −t0 )
VT (x) VT (b)
µ̂ (x | b) = −
Vt0 (x) Vt0 (b)
1 ∑
T
ŝ (x | b) = st (x | b)
n t=t
0

1 ∑T
1 ∑ T
µ̄ (x | b) = Rt (x) − Rt (b)
n − 1 t=t +1 n − 1 t=t +1
v
0 0

u
u 260 ∑ T
= t
2
σ̂ (x | b) (Rt (x) − Rt (b) − µ̄ (x | b))
n − 1 t=t +1
0

where Vt (x) (or Vt (b)) is the net asset value of the tracker (or benchmark) at time t, st (x | b)
is the bid-ask spread9 at time t and Rt (x) (or Rt (b)) is the daily return of the tracker (or
the benchmark) defined as follows:
Vt (x)
Rt (x) = −1
Vt−1 (x)
9 More precisely, we compute the best spread of the first limit order for each listing place and each trading

day t. st (x | b) is then the weighted average by considering the daily volume of the different listing places.

14
Measuring Performance of Exchange Traded Funds

3.1 An application to European ETFs


The study period begins in November, 30th 2011 and ends in November, 30th 2012. We
consider three benchmarks: the Eurostoxx 50 index, the S&P 500 index and the MSCI
World index. To compute the statistic ζα (x | b) properly, we need to make some statistical
adjustments:
• Some ETFs distribute dividends. In this case, we have to rebuild the net asset value by
incorporating these dividends in order to compute performance that can be compared
to the performance of the benchmark.
• Dividends also influence the tracking-error volatility. That is why we correct the daily
performance of the tracker every time a dividend is paid.
• Some ETFs have poor liquidity meaning, that we cannot observe a spread for each
trading date. To compute ŝ (x | b), we decide to exclude dates with zero trading
volume.
• ETFs may be listed in different currencies. We therefore adopt the point of view of a
European investor and consider the Euro as the default currency.

Table 1: Computation of the efficiency measure (Eurostoxx 50)

Tracker µ̂ (x) µ̂ (x | b) ŝ (x | b) σ̂ (x) σ̂ (x | b) ζα (x | b)


Amundi 15.29 62.56 9.84 22.33 11.97 32.98
db X-trackers 15.33 65.97 12.27 22.86 7.31 41.64
iShares (DE) 15.05 37.88 7.96 21.62 56.54 −63.38
iShares 15.25 58.46 10.39 21.92 19.62 15.70
Lyxor 15.30 63.51 8.48 22.01 14.89 30.47
Source 14.90 23.51 15.38 22.23 7.25 −3.83
Index 14.67 22.09

Table 2: Computation of the efficiency measure (S&P 500)

Tracker µ̂ (x) µ̂ (x | b) ŝ (x | b) σ̂ (x) σ̂ (x | b) ζα (x | b)


Amundi 19.49 9.19 16.97 12.59 3.14 −12.97
Credit Suisse 19.57 16.99 18.23 12.50 4.63 −8.88
db X-trackers 19.56 16.04 18.26 12.79 4.65 −9.90
HSBC 19.68 28.20 20.58 12.68 3.45 1.92
iShares 19.34 −6.10 7.45 12.56 4.90 −21.63
Lyxor 19.60 19.87 13.56 12.59 0.98 4.69
Source 19.34 −5.30 17.81 12.87 1.78 −26.04
UBS 19.40 0.02 41.13 12.85 0.59 −42.09
Index 19.40 12.76

Results10 are reported in Tables 1, 2 and 3. According to the efficiency measure ζα (x | b),
the best tracker for the Eurostoxx 50 index is db X-trackers, followed by Amundi and Lyxor.
We notice that ζα (x | b) assigns a positive value to most of the trackers, due probably to
10 All the statistics are expressed in bps except µ̂ (x) and σ̂ (x) which are measured in %.

15
Measuring Performance of Exchange Traded Funds

fiscal optimization of dividends. These results are different for the S&P 500 and MSCI World
indices. Indeed, ζα (x | b) is generally negative. It is due to lower performance and a wider
spread, even if ETFs tracking the S&P 500 and MSCI World indices have less tracking-error
volatility11 than for the Eurostoxx 50 index. At the end, Lyxor is ranked top, followed by
HSBC and Credit Suisse for the S&P 500 index, and Commerzbank and Amundi for the
MSCI World index.

Table 3: Computation of the efficiency measure (MSCI World)

Tracker µ̂ (x) µ̂ (x | b) ŝ (x | b) σ̂ (x) σ̂ (x | b) ζα (x | b)


Amundi 17.40 −20.75 23.91 10.32 3.67 −50.72
Commerzbank 17.42 −18.25 18.72 10.83 3.56 −42.85
db X-trackers 17.36 −24.46 13.20 10.30 25.18 −79.20
iShares 17.18 −42.08 13.75 10.14 50.80 −139.65
Lyxor 17.37 −23.09 11.93 10.09 1.55 −37.58
Source 17.08 −52.64 24.83 10.31 1.69 −80.25
UBS 17.36 −23.98 31.25 10.39 14.51 −79.16
Index 17.60 10.18

Table 4: Computation of the efficiency measure (MSCI EM)

Tracker µ̂ (x) µ̂ (x | b) ŝ (x | b) σ̂ (x) σ̂ (x | b) ζα (x | b)


Credit Suisse 13.20 −205.79 30.05 13.18 150.68 −484.46
db X-trackers 14.13 −112.34 15.85 13.07 12.83 −149.35
iShares 14.18 −107.56 17.90 13.07 160.21 −389.80
Lyxor 14.45 −80.01 20.72 13.07 14.87 −125.28
Source 14.16 −109.20 50.30 13.22 3.96 −166.02
Index 15.25 13.12

3.2 Different benchmarks


One difficulty arises when ETF providers do not choose the same benchmark to give access to
an asset class. For instance, if we consider Japanese equities, we notice that some providers
have chosen to replicate the Topix index, whereas other providers have opted for the MSCI
Japan index. In this case, results depend on the reference index. Nevertheless, although the
two benchmarks are highly correlated, they may be coherent and the choice of the reference
index has a limited impact on the ranking. In this situation, we use principal component
analysis to build a reference index, as suggested by Amenc and Martellini (2002). The
idea is to extract the best possible one-dimensional summary of the set of these different
benchmarks.

For instance, in our example, the reference index is composed by 49.50% of the Topix
index and 50.50% of the MSCI Japan index12 , and we obtain the results given in Table 5.
For Japanese equities, iShares appears then the best tracker.
11 Nevertheless, we note some divergence in the case of MSCI World trackers. This is even more so when

we consider MSCI EM trackers (see Table 4.


12 Methodological details to compute these weights are provided in Appendix A.3.

16
Measuring Performance of Exchange Traded Funds

Table 5: Computation of the efficiency measure (Japanese equities)

Tracker µ̂ (x) µ̂ (x | b) ŝ (x | b) σ̂ (x) σ̂ (x | b) ζα (x | b)


db X-trackers −0.72 −35.87 15.82 16.58 256.73 −475.29
iShares −0.76 −40.64 19.98 15.97 66.84 −170.90
Lyxor −1.29 −92.96 14.98 16.10 62.81 −211.58
Source −0.67 −31.40 35.11 16.58 65.30 −174.25
Index −0.36 15.84

4 Variations on the tracker efficiency measure


4.1 Tracking error volatility or semi-volatility?
One could argue that tracking error volatility is not a reliable measure of risk, because
it depends on both positive and negative tracking errors. That is why different authors
suggest using semi-variance instead of variance (Markowitz, 1959; Fishburn, 1977; Harlow,
1991; Estrada, 2007). Semi-variance is a special case of lower partial moments (see Appendix
A.4). In Table 7, we report the values taken by the semi-volatility σ̂− (x | b) of the tracking
error for the Eurostoxx 50 trackers. We consider two values for the threshold. When
τ = µ (x | b), the semi-volatility σ̂− (x | b) corresponds to the square root of the below-mean
semi-variance. When τ = 0, the semi-volatility σ̂− (x | b) is the square root of below-target
semi-variance (Nawrocki, 1999). By definition, semi-volatility is lower than volatility. We
verify this property in Table 7. If the distribution of tracking errors is perfectly√ symmetric
around the mean, the ratio between volatility and semi-volatility is equal to 2. That is
why we compute the modified efficiency measure as follows:

ζα⋆ (x | b) = µ̂ (x | b) − ŝ (x | b) − 1.65 · 2 · σ̂− (x | b)

We notice that the ranking changes slightly in comparison to the previous one. Amundi is
thus the best tracker, followed by db X-trackers and Lyxor.

Table 6: Estimated risk of Eurostoxx 50 trackers


τ = µ (x | b) τ =0
Tracker σ̂ (x | b) ζα (x | b) σ̂− (x | b) ζα⋆ (x | b) σ̂− (x | b) ζα⋆ (x | b)
Amundi 11.97 32.98 3.43 44.72 1.05 50.28
db X-trackers 7.31 41.64 4.11 44.12 2.86 47.04
iShares (DE) 56.54 −63.38 40.38 −64.32 40.01 −63.45
iShares 19.62 15.70 11.86 20.38 11.00 22.40
Lyxor 14.89 30.47 5.74 41.63 4.28 45.05
Source 7.25 −3.83 1.88 3.74 1.06 5.66

Looking at the results in Table 7, the ranking of S&P 500 trackers is unchanged: Lyxor
remains the best tracker, followed by HSBC and Credit Suisse. We conclude that the choice
of semi-variance over volatility does not fundamentally alter the results. But the efficiency
measure based on semi-variance has an important drawback. It is less sensitive to tracking
errors, which is however an important criteria for investors. Another important point is that
semi-variance is just another way of measuring the skewness of a distribution:

17
Measuring Performance of Exchange Traded Funds

Table 7: Estimated risk of S&P 500 trackers


τ = µ (x | b) τ =0
Tracker σ̂ (x | b) ζα (x | b) σ̂− (x | b) ζα⋆ (x | b) σ̂− (x | b) ζα⋆ (x | b)
Amundi 3.14 −12.97 1.22 −10.64 0.84 −9.74
Credit Suisse 4.63 −8.88 3.01 −8.26 2.53 −7.13
db X-trackers 4.65 −9.90 3.39 −10.13 2.91 −9.01
HSBC 3.45 1.92 2.01 2.92 1.25 4.71
iShares 4.90 −21.63 3.26 −21.15 3.34 −21.34
Lyxor 0.98 4.69 0.68 4.73 0.28 5.67
Source 1.78 −26.04 1.10 −25.67 1.23 −25.97
UBS 0.59 −42.09 0.37 −41.98 0.35 −41.93

“By taking the variance and dividing it by the below-mean semi-variance, a


measure of skewness resulted. If the distribution is normally distributed then
the semi-variance should be one-half of the variance [...] If the ratio is not
equal to 2, then there is evidence that the distribution is skewed or asymmetric”
(Nawrocki, 1999, page 11).

If we compute this ratio for the previous trackers, we get a range between 1.9 and 14.8
for the Eurostoxx 50 index and between 1.9 and 6.6 for the S&P 500 index. This clearly
demonstrates that tracking errors exhibit high skewness. This result is confirmed by the
Kernel analysis of empirical density (see Figures 9, 10 and 11). In this context, we think
that replacing volatility by semi-volatility is not a good method. It would be better to adopt
another risk measure that takes the skewness of tracking errors into account.

Figure 9: Estimated density of daily tracking errors (Eurostoxx 50)

18
Measuring Performance of Exchange Traded Funds

Figure 10: Estimated density of daily tracking errors (S&P 500)

Figure 11: Estimated density of daily tracking errors (S&P 500)

19
Measuring Performance of Exchange Traded Funds

4.2 Choosing another risk measure


The efficiency measure ζα (x | b) is based on the assumption that tracking errors are normally
distributed. In this paragraph, we explore non-gaussian risk measures. For instance, we can
replace gaussian value-at-risk by historical value-at-risk or the expected shortfall. One of
the difficulties with these two risk measures is that they are generally computed on a daily
basis, but the conversion to an annual measure is not straightforward. In the previous
gaussian value-at-risk, we adopt the square root rule: annual volatility of tracking errors is
obtained
√ by considering the product of the daily volatility of tracking errors and the factor
260. This rule is universally accepted and has been used in fund reports for many years.
Unfortunately, the square root of time is appropriate when returns are normally distributed,
but may be not so in other cases (McNeil and Frey, 2000). Because there is no solution to
this problem, professionals still use the square root rule in the non-gaussian case, but only
when returns are centered. Following this common practice, we suggest defining the efficient
indicator as follows:
ζα⋆ (x | b) = µ (x | b) − s (x | b) − F−1
0 (α)
where F0 is the probability distribution of centered tracking errors.

In Tables 8 and 9, we compute the efficiency indicator ζα (x | b) with five risk measures:
σ (x | b) corresponds to volatility, σ− (x | b) is below-mean semi-volatility, VaRα (x | b) is
historical value-at-risk, ESα (x | b) corresponds to the historical expected shortfall, whereas
CFα (x | b) uses the Cornish-Fisher expansion to compute value-at-risk. In the case of Eu-
rostoxx 50 trackers, we observe some minor differences. Indeed, Lyxor turns out to be the
best tracker if we are looking at historical value-at-risk. The rank of db X-trackers varies
between 1 and 3 depending on the risk measures. For S&P 500 trackers, we observe a good
consistency across the various rankings.

Table 8: Efficiency measures of Eurostoxx 50 trackers

Tracker σ (x | b) σ− (x | b) VaRα (x | b) ESα (x | b) CFα (x | b)


Amundi 32.98 44.72 47.75 47.25 62.54
db X-trackers 41.64 44.12 45.90 41.02 47.15
iShares (DE) −63.38 −64.32 −59.40 −102.51 −62.72
iShares 15.70 20.38 34.77 12.21 27.93
Lyxor 30.47 41.63 47.97 38.26 61.48
Source −3.83 3.74 5.00 4.63 19.17

Table 9: Efficiency measures of S&P 500 trackers

Tracker σ (x | b) σ− (x | b) VaRα (x | b) ESα (x | b) CFα (x | b)


Amundi −12.97 −10.64 −10.24 −10.55 −8.04
Credit Suisse −8.88 −8.26 −6.60 −10.37 −7.35
db X-trackers −9.90 −10.13 −8.88 −11.77 −10.02
HSBC 1.92 2.92 4.24 2.23 3.73
iShares −21.63 −21.15 −19.10 −23.59 −20.23
Lyxor 4.69 4.73 4.88 4.25 4.71
Source −26.04 −25.67 −25.50 −26.00 −25.37
UBS −42.09 −41.98 −41.93 −42.09 −41.95

20
Measuring Performance of Exchange Traded Funds

4.3 Taking the liquidity risk into account


The spread measure s (x | b) used previously corresponds to the daily average of the first
limit order spreads. This measure may be pertinent for a retail investor, but is not for an
institutional investor. Institutional investors buy or sell a notional N , that can not generally
be executed via the best first limit orders. That is why we consider another spread measure
sN (x | b) corresponding to intraday spreads weighted by the duration between two ticks for
a given notional13 . To illustrate the use of this spread, we are going to look at the Eurostoxx
50 trackers.

Figure 12: Evolution of the spread of the Amundi tracker

In Figure 12, we shown the change in the spread sN (x | b) for the Amundi Eurostoxx
50 tracker and different values of the notional N . We notice that this spread changes with
respect to market liquidity. For each tracker and each notional value, we have shown the
corresponding boxplot in Figure 13. The boxplot indicates the minimum value, the quartile
range, the median and the last decile. We check that:

N1 ≥ N2 ⇒ sN1 (x | b) ≥ sN2 (x | b)

We notice also that for some trackers, the spread increases greatly in line with volume.
In Figure 14, we have drawn some scatterplots. It is interesting to note that when liquidity
is not an issue, the trackers are more or less equivalent in terms of spread14 . This is not the
case when there are some liquidity problems (see Figure 15).
13 Computational details are provided in Appendix A.5.
14 If we consider two trackers x and y, we have:
sN (x | b) ≃ sN (y | b)
if sN (x | b) ≃ 0 and sN (x | b) ≃ 0.

21
Measuring Performance of Exchange Traded Funds

Figure 13: Boxplot of ETF spreads

Figure 14: Scatterplot of ETF spreads with a volume of 1 MEUR

22
Measuring Performance of Exchange Traded Funds

Figure 15: Quantile regression of ETF spreads with a volume of 1 MEUR

Let us now compute the efficiency indicator ζα⋆ (x | b) by considering this new spread
measure. The results are shown in Table 10. We consider two calculations: one based on
the median spread and another one based on the quantile at the 95% confidence level. We
notice that this new measure has a big impact on the efficiency indicator. For instance, if
we consider a notional of 2 MEUR, we obtain a different ranking.

Table 10: Impact of the liquidity on the efficiency measure (Eurostoxx 50)

100 KEUR 1 MEUR 2 MEUR


Tracker
50% 95% 50% 95% 50% 95%
Amundi 35.01 30.19 32.96 25.45 28.52 15.00
db X-trackers 45.48 43.28 41.65 −2.39 37.29 −66.68
iShares (DE) −62.66 −65.05 −65.00 −77.04 −66.74 −98.73
iShares 18.30 15.72 15.85 10.33 13.04 5.31
Lyxor 31.93 27.89 29.39 24.98 27.23 19.97
Source −1.46 −8.80 −9.80 −64.29 −101.00 −193.50

Remark 2 In our definition of the efficiency indicator, we assume that there is one trade
each year. If there are m trades per year, the performance measure becomes:
ζα (x | b) = µ (x | b) − m · s (x | b) − Φ−1 (α) σ (x | b)
This formula highlights the importance of liquidity for active managers. For instance, a
highly active manager will only be interested in the spread measure because:
lim ζα (x | b) = −m · s (x | b)
m→∞

23
Measuring Performance of Exchange Traded Funds

5 Conclusion
In this paper, we have developed a performance measure to compare passive management,
in particular tracker investment vehicles such as exchange traded funds. This measure is
very different from the ones used to assess the performance of active management. It is a
value-at-risk measure based on three parameters: the performance difference between the
fund and the index, the volatility of the tracking error and the liquidity spread. This simple
measure may be easily implemented by investors or rating agencies in relation to the fund
picking process.

This paper also highlights the role of liquidity spread in measuring the efficiency of an
ETF. It is particular true for institutional investors, who may subscribe or redeem large
investment amounts. The liquidity spread is also the most important parameter for active
managers who use ETFs to implement their convictions in line with their tactical asset
allocation.

24
Measuring Performance of Exchange Traded Funds

References
[1] Amenc N. and Martellini L. (2002), The Brave New World of Hedge Fund Indices,
EDHEC-Risk Working paper.

[2] Barras L., Scaillet O. and Wermers R. (2010), False Discoveries in Mutual Fund
Performance: Measuring Luck in Estimated Alphas, Journal of Finance, 65(1), pp.
179-216.

[3] Bawa V.S. (1975), Optimal Rules for Ordering Uncertain Prospects, Journal of Finan-
cial Economics, 2(1), pp. 95-121.

[4] Beasley J.E., Meade N. and Chang T.-J. (2003), An Evolutionary Heuristic for
the Index Tracking Problem, European Journal of Operational Research, 148(3), pp.
621-643.

[5] Blake C.R., Elton E.J. and Gruber M.J. (1993), The Performance of Bond Mutual
Funds, Journal of Business, 66(3), pp. 371-403.

[6] Bernstein P.L. (1992), Capital Ideas: The Improbable Origins of Modern Wall Street,
Free Press.

[7] Bernstein P.L. (2007), Capital Ideas Evolving, John Wiley & Sons.

[8] Carhart M.M. (1997), On Persistence in Mutual Fund Performance, Journal of Fi-
nance, 52(1), pp. 57-82.

[9] Dieterlen R. and Hereil P. (2012), Index Arbitrage Explained, Journal of Indexes
Europe, 2(6), pp. 18-23.

[10] Elton E.J., Gruber M.J., Das S. and Hlavka M. (1993), Efficiency with Costly
Information: A Reinterpretation of Evidence from Managed Portfolios, Review of Fi-
nancial Studies, 6(1), pp. 1-22.

[11] Elton E.J., Gruber M.J. and Busse J.A. (2004), Are Investors Rational? Choices
among Index Funds, Journal of Finance, 59(1), pp. 261-288.

[12] Estrada J. (2007), Mean-semivariance Behavior: Downside Risk and Capital Asset
Pricing, International Review of Economics and Finance, 16(2), pp. 169-185.

[13] Fishburn P.C. (1977), Mean-Risk Analysis with Risk Associated with Below-Target
Returns, American Economic Review, 67(1), pp. 116-126.

[14] Gastineau G.L. (2002), Equity Index Funds Have Lost Their Way, Journal of Portfolio
Management, 28(2), pp. 55-64.

[15] Gastineau G.L. (2004), The Benchmark Index ETF Performance Problem, Journal
of Portfolio Management, 30(2), pp. 96-103.

[16] Grinblatt M. and Titman S. (1989), Portfolio Performance Evaluation: Old Issues
and New Insights, Review of Financial Studies, 2(3), pp. 393-421.

[17] Grinold R.C. and Kahn R.N. (2000), Active Portfolio Management: A Quantita-
tive Approach for Providing Superior Returns and Controlling Risk, Second edition,
McGraw-Hill.

25
Measuring Performance of Exchange Traded Funds

[18] Harlow W.V. (1991), Asset Allocation in a Downside-Risk Framework, Financial


Analysts Journal, 47(5), pp. 28-40.
[19] Jensen M.C. (1968), The Performance of Mutual Funds in the Period 1945-1964, Jour-
nal of Finance, 23(2), pp. 389-416.
[20] Jorion P. (2003), Portfolio Optimization with Tracking-Error Constraints, Financial
Analysts Journal, 59(5), pp. 70-82.
[21] Markowitz H. (1952), Portfolio Selection, Journal of Finance, 7(1), pp. 77-91.

[22] Markowitz H. (1959), Portfolio Selection: Efficient Diversification of Investments,


John Wiley & Sons.
[23] McNeil A.J. and Frey R. (2000), Estimation of Tail-related Risk Measures for Het-
eroscedastic Financial Time Series: An Extreme Value Approach, Journal of Empirical
Finance, 7(3-4), pp. 271-300.

[24] Nawrocki D.N. (1999), A Brief History of Downside Risk Measures, Journal of In-
vesting, 8(3), pp. 9-25.

[25] Pope P.F. and Yadav P.K. (1994), Discovering Errors in Tracking Error, Journal of
Portfolio Management, 20(2), pp. 27-32.

[26] Prigent J.-L. (2007), Portfolio Optimization and Performance Analysis, Chapman &
Hall.

[27] Roll R. (1992), A Mean/Variance Analysis of Tracking Error, Journal of Portfolio


Management, 18(4), pp. 13-22.

[28] Sharpe W.F. (1964), Capital Asset Prices: A Theory of Market Equilibrium under
Conditions of Risk, Journal of Finance, 19(3), pp. 425-442.

[29] Tobin J. (1958), Liquidity Preference as Behavior Towards Risk, Review of Economic
Studies, 25(2), pp. 65-86.

26
Measuring Performance of Exchange Traded Funds

A Technical appendix
A.1 Proof of the equation (1)
We have:

σ 2 (x | b) = (x − b) Σ (x − b)
= x⊤ Σx + b⊤ Σb − 2x⊤ Σb
= σ 2 (x) + σ 2 (b) − 2ρ (x, b) σ (x) σ (b)

We deduce that the correlation between the portfolio x and the benchmark b is:

σ 2 (x) + σ 2 (b) − σ 2 (x | b)
ρ (x, b) =
2σ (x) σ (b)
It follows that:
⊤ ( )
2 (x0 − b) Σ (x − b) = 2 x⊤ ⊤ ⊤ ⊤
0 Σx − x0 Σb − x Σb + b Σb
= σ 2 (x0 ) + σ 2 (x) − σ 2 (x | x0 ) −
σ 2 (x0 ) − σ 2 (b) + σ 2 (x0 | b) −
σ 2 (x) − σ 2 (b) + 2σ 2 (x | b) + σ 2 (b)
= σ 2 (x0 | b) + σ 2 (x | b) − σ 2 (x | x0 )

The weights of the portfolio y are given by:

y = (1 − α) x0 + αx

We deduce that:
y − b = (1 − α) (x0 − b) + α (x − b)
It follows that:
µ (y | b) = (1 − α) µ (x0 | b) + αµ (x | b)
and:

σ 2 (y | b) = (y − b) Σ (y − b)
2 ⊤
= (1 − α) (x0 − b) Σ (x0 − b) +

α2 (x − b) Σ (x − b) +

2α (1 − α) (x0 − b) Σ (x − b)
2
= (1 − α) σ 2 (x0 | b) + α2 σ 2 (x | b) +
( )
α (1 − α) σ 2 (x0 | b) + σ 2 (x | b) − σ 2 (x | x0 )
( )
= (1 − α) σ 2 (x0 | b) + ασ 2 (x | b) + α2 − α σ 2 (x | x0 )

We deduce that:
µ (y | b)
IR (y | b) =
σ (y | b)
(1 − α) µ (x0 | b) + αµ (x | b)
= √
(1 − α) σ (x0 | b) + ασ 2 (x | b) + (α2 − α) σ 2 (x | x0 )
2

27
Measuring Performance of Exchange Traded Funds

A.2 Derivation of the efficiency measure ζα (x | b)


The probability distribution function of the tracking error e is Gaussian with:
( )
e ∼ N µ (x | b) , σ 2 (x | b)

We have:

Pr {L (x | b) ≤ ζ} = α
⇔ Pr {e − s ≤ ζ} = α
⇔ Pr {e ≤ s + ζ} = α
{ }
e − µ (x | b) s + ζ − µ (x | b)
⇔ Pr ≤ =α
σ (x | b) σ (x | b)
( )
s + ζ − µ (x | b)
⇔ Φ =α
σ (x | b)
It follows that:
s + ζ − µ (x | b)
= Φ−1 (α)
σ (x | b)
or:
ζ = s − µ (x | b) + Φ−1 (α) σ (x | b)
We deduce that:

ζα (x | b) = −ζ
= µ (x | b) − s − Φ−1 (α) σ (x | b)

A.3 Building a benchmark from a set of competing indices


Let {b1 , . . . , bm } be a set of competing indices. The underlying idea is to compute a bench-
mark b representative of these indices. Let Ω be the m × m covariance matrix of the returns
Rt (bj ). By computing the eigendecomposition of Ω, we get:

Ω = V ΛV ⊤ (3)

where V is the matrix of the eigenvectors and Λ = diag (λ1 , . . . , λm ) is a diagonal matrix
of eigenvalues15 . Because the decomposition (3) corresponds to the principal component
analysis (PCA), we may extract the main representative factor defined by:

m
Ft = Vj,1 Rt (bj )
j=1

As noticed by Amenc and Martellini (2002), this first component of the PCA maximizes the
representativeness of the set of competing indices. The reference index is then a weighted
portfolio of the competing indices:
∑m
b= wj bj
j=1

where the weights are proportional to the loading coefficients:


Vj,1
wj = ∑m
k=1 Vk,1
15 We assume that λ1 ≥ . . . ≥ λm .

28
Measuring Performance of Exchange Traded Funds

A.4 Lower partial moments


Let X be a random variable with distribution F. The mathematical expectation of X is16 :

µ = E [X]
∫ ∞
= x dF (x)
−∞
∫ ∞
= xf (x) dx
−∞

whereas the centered moment of order n is:


n
µn (X) = E [(X − µ) ]
∫ ∞
n
= (x − µ) f (x) dx
−∞

The variance of X corresponds to the second moment: σ 2 (X) = µ2 (X). σ (X) is called
standard deviation, but it is well known as volatility in finance, where X represents the
return on an asset.

The lower partial moment (LPM) is obtained by (Bawa, 1975):


n
LPMn (X; τ ) = E [max (0, τ − X) ]
∫ τ
n
= (τ − x) f (x) dx
−∞

where τ is a threshold. If τ is equal to the mean µ, we get:


n
LPMn (X; µ) = E [max (0, µ − X) ]
∫ µ
n
= (µ − x) f (x) dx
−∞

Semi-variance is then defined as the second lower partial moment:

SV (x) = LPM2 (X; µ)


[ ]
2
= E max (0, µ − X)
∫ µ
2
= (µ − x) f (x) dx
−∞

If the distribution of X is symmetric around the mean, we get:


∫ ∞ ∫ µ ∫ ∞
2 2 2
(µ − x) f (x) dx = (µ − x) f (x) dx + (µ − x) f (x) dx
−∞ −∞ µ
∫ µ
2
= 2 (µ − x) f (x) dx
−∞

We deduce that semi-variance is half of the variance.


16 We note f (x) the associated density function.

29
Measuring Performance of Exchange Traded Funds

A.5 Computation of the spread sN (x | b)


A.5.1 Analytical expression
We define the daily spread sN (x | b) as a weighted average of intraday spreads:
∑close
j=open sj (tj+1 − tj )
sN (x | b) = ∑close
j=open (tj+1 − tj )

where sj is the spread of the j th tick and tj+1 − tj the elapsed time between two consecutive
ticks: ( + )
P̄j − P̄j−
sj = cj
P̄j0
We have also: ∑K
Q̄•j,k Pj,k

P̄j•
k=1
= ∑K
k=1 Q̄•j,k
+ −
where Pj,k (resp. Pj,k ) is the ask (or bid) price at tj for the k th limit order. The average
0
mid price P̄j corresponds to:
P̄j+ + P̄j−
P̄j0 =
2

The quantity Q̄+
j,k and Q̄j,k are defined as follows:
( ( ))

k−1
Q̄•j,k = max 0, min Q•j,k , Q⋆j − Q•j,l
l=1


Here, Q+j,k and Qj,k are the ask and bid volumes of the k
th
limit order. The reference

quantity Qj is the ratio between the trading notional N and the mid price:

N
Q⋆j =
P̄j0

Sometimes it may appear that the trading volume on the order book is lower than the
notional N . That is why the factor cj may be greater than one:
 
Q⋆j
cj = max 1, (∑ )
K + ∑K −
min k=1 Qj,k , k=1 Qj,k

For instance, if we wish to execute an order of 2 MEUR and there is only a trading volume
of 1 MEUR, we multiply the spread by two.

Remark 3 For each trading day, we compute the daily spread for the different listing places
using the previous formulas and we take the best spread.

A.5.2 Example
Let us illustrate the spread calculation with the order book given in Table 11. For Q⋆ = 1000,
we get the results shown in Table 12. We deduce that the mid price P̄j0 is 85.98 whereas the
spread sj is equal to 20.12 bps. It corresponds to a notional N = Q⋆ × P̄j0 of 85 981 e.

30
Measuring Performance of Exchange Traded Funds

Table 11: Limit order book


Buy orders Sell orders
k
Q−j,k

Pj,k Q+j,k
+
Pj,k
1 900 85.90 600 86.05
2 200 85.85 300 86.06
3 57 85.82 400 86.20
4 18 85.75 213 86.21
5 117 85.74 73 86.22
6 1000 85.73 200 86.23
7 3000 85.72 1500 86.25

Table 12: Computing the spread for Q⋆ = 1000

Buy orders Sell orders


k
Q̄−
j,k

Pj,k Q̄+
j,k
+
Pj,k
1 900 85.90 600 86.05
2 100 85.85 300 86.06
3 0 85.82 100 86.20
∑K •
k=1 Q̄j,k 1000 1000

P̄j 85.89 86.07

If we wish to execute an order for 100 KEUR, it implies that Q⋆ > 1000. Given a notional
N , it is then possible to find the optimal value of Q⋆ :
{ }
Q⋆ = inf Q ∈ N : QP̄j0 ≥ N

We may solve this nonlinear inequality by using a bisection algorithm. For instance, if N is
equal to 100 KEUR, we obtain Q⋆ = 1163. In this case, the spread is equal to 23.24 bps.
Sometimes, there are not enough orders in the book. For instance, if N = 500 KEUR, the
mid price is 85.972 meaning that Q⋆ = 5816. The coefficient c is then larger than 1 and we
obtain a spread equal to 87.81 bps.

Table 13: Computing the spread for a given notional N

N = 100 KEUR N = 500 KEUR


k Buy orders Sell orders Buy orders Sell orders
Q̄−
j,k

Pj,k Q̄+
j,k
+
Pj,k Q̄−
j,k

Pj,k Q̄+
j,k
+
Pj,k
1 900 85.90 600 86.05 900 85.90 600 86.05
2 100 85.85 300 86.06 200 85.85 300 86.06
3 57 85.82 263 86.20 57 85.82 400 86.20
4 6 85.75 0 86.21 18 85.75 213 86.21
5 0 85.74 0 86.22 117 85.74 73 86.21
6 0 85.73 0 86.23 1000 85.73 200 86.23
7 0 85.72 0 86.25 3000 85.72 1500 86.25
∑K •
k=1 Q̄j,k 1163 1163 5292 3286

P̄j 85.89 86.09 85.76 86.19

31
Measuring Performance of Exchange Traded Funds

B Description of the data


In the following table, for each universe we show the name of the tracker, its description and
the corresponding Bloomberg code.

Universe Name Description BBG code

Amundi AMUNDI ETF DJE 50 C50 FP


db X-trackers DBX-TRACKERS EURO STXX 50 XESX GY
iShares (DE) ISHARES EURO STOXX 50 DE SX5EEX GY
Eurostoxx 50 iShares ISHARES EURO STOXX 50 EUN2 GY
Lyxor LYXOR ETF Euro Stoxx 50 MSE FP
Source EURO STOXX 50 Source ETF SDJE50 GY
Eurostoxx 50 ESTX 50 e NRt SX5T

Amundi AMUNDI ETF S&P 500 500 FP


Credit Suisse CSETF ON S&P 500 CSSPX SW
db X-trackers DB X-TRACKERS S&P 500 ETF D5BM GY
HSBC HSBC S&P 500 ETF HSPD LN
S&P 500 iShares ISHARES S&P 500 Index Fund IUSA LN
Lyxor LYXOR ETF S&P 500 SP5 FP
Source S&P 500 SOURCE ETF SPXS LN
UBS UBS S&P 500 INDEX-A S5USAS SW
S&P 500 S&P 500 Net TR SPTR500N

Amundi AMUNDI ETF MSCI WORLD CW8 FP


Commerzbank CCOMSTAGE ETF MSCI WORLD-I CBNDDUWI GY
db X-trackers DB X-TRACKERS MSCI WORLD TRN XMWO GY
iShares ISHARES MSCI World IWRD LN
MSCI World
Lyxor LYXOR ETF MSCI World WLD FP
Source MSCI-WORLD SOURCE ETF SMSWLD GY
UBS UBS-ETF MSCI WORLD WRDCHA SW
MSCI World MSCI Daily TR Net World USD NDDUWI

Credit Suisse CS ETF ON MSCI EMERGING MRKT CSEM SW


db X-trackers DB X-TRACKERS EMERG MARKET XMEM GY
iShares iShares MSCI Emerging Markets (EUR) IEMM NA
MSCI EM
Lyxor LYXOR ETF MSCI Emerging Markets LEM FP
Source MSCI EMERG MARKET SOURCE ETF MXFS LN
MSCI EM MSCI Daily TR Net Emerging Markets NDUEEGF

db X-trackers DB X-TRACKERS MSCI JAPAN TRN XMJP GY


iShares ISHARES MSCI JAPAN FUND IJPN LN
Lyxor LYXOR ETF Japan (TOPIX) JPN FP
Japanese Equities
Source MSCI JAPAN SOURCE ETF SMSJPN GY
Topix TOPIX TR Index TPXDDVD
MSCI Japan MSCI Daily TR Gross Japan USD GDDUJN

32

Vous aimerez peut-être aussi